<<

1 2 3 4 5 Smectite formation in the presence of sulfuric acid: Implications for

6 acidic smectite formation on early

7 8 9 10 11 T.S. Peretyazhko1*, P.B. Niles2, B. Sutter1, R. V. Morris2, D. G. Agresti3, L. Le1, D.W. Ming2 12 13 14 15 16 17 18 1 Jacobs, NASA Johnson Space Center, Houston, TX 77058 19 20 2 NASA Johnson Space Center, Houston, TX 77058 21 3 University of Alabama at Birmingham, Birmingham, AL 35294 22 23 24 25 26 27 Corresponding author: T. S. Peretyazhko, Jacobs, NASA Johnson Space Center, 28 Houston, TX 77058 ([email protected]) 29 30 31 32 33 In preparation for

34 Geochimica et Cosmochimica Acta 35 Abstract

36 The excess of orbital detection of smectite deposits compared to carbonate deposits on the

37 surface presents an enigma because smectite and carbonate formations are both favored alteration

38 products of under neutral to alkaline conditions. We propose that Mars experienced acidic

39 events caused by sulfuric acid (H2SO4) that permitted phyllosilicate, but inhibited carbonate,

40 formation. To experimentally verify this hypothesis, we report the first synthesis of smectite from

41 Mars-analogue glass-rich basalt simulant (66 wt% glass, 32 wt% olivine, 2 wt% chromite) in the

42 presence of H2SO4 under hydrothermal conditions (~200 °C). Smectites were analyzed by X-ray

43 diffraction, Mössbauer spectroscopy, visible and near-infrared reflectance spectroscopy and

44 electron microprobe to characterize mineralogy and chemical composition. Solution chemistry was

45 determined by Inductively Coupled Plasma Mass Spectrometry. Basalt simulant suspensions in 11-

46 42 mM H2SO4 were acidic with pH ≤ 2 at the beginning of incubation and varied from acidic (pH

47 1.8) to mildly alkaline (pH 8.4) at the end of incubation. Alteration of glass phase during reaction

48 of the basalt simulant with H2SO4 led to formation of the dioctahedral smectite at final pH ~3 and

49 trioctahedral smectite saponite at final pH ~4 and higher. Anhydrite and hematite formed in the

50 final pH range from 1.8 to 8.4 while natroalunite was detected at pH 1.8. Hematite was

51 precipitated as a result of oxidative dissolution of olivine present in Adirondack basalt simulant.

52 Formation of secondary phases, including smectite, resulted in release of variable amounts of Si,

53 Mg, Na and Ca while solubilization of Al and Fe was low. Comparison of mineralogical and

54 solution chemistry data indicated that the type of smectite (i.e., dioctahedral vs trioctahedral) was

55 likely controlled by Mg leaching from altering basalt and substantial Mg loss created favorable

56 conditions for formation of dioctahedral smectite. We present a model for global-scale smectite

57 formation on Mars via acid-sulfate conditions created by the volcanic outgassing of SO2 in the

58 and early .

2

59 1. INTRODUCTION

60

61 Two global eras have been proposed to explain the observed mineralogy on Mars

62 (Bibring et al., 2006). Abundant phyllosilicates dominated by the smectite group (nontronite,

63 montmorillonite and saponite) were formed in the first era under water-rich neutral to alkaline

64 conditions during the Noachian. Formation of sulfate-bearing phases occurred in the second era

65 under acidic conditions likely caused by sulfuric acid during the Hesperian (Poulet et al., 2005;

66 Bibring et al., 2006; Murchie et al., 2009; Bishop et al., 2013). However, such simplified pH-

67 based division of aqueous history conflicts with some mineralogical observations. Large

68 carbonate deposits together with phyllosilicates would be a characteristic of a Noachian Mars

69 dominated by abundant liquid water, neutral/alkaline pH conditions, and a CO2-rich atmosphere

70 (Fairén et al., 2004), but mineralogical observations have detected only isolated carbonate

71 deposits (Ehlmann et al., 2008; Milliken et al., 2009; Morris et al., 2010; Wray et al., 2016). The

72 absence of widespread carbonate deposits could result from carbonate deposition on early Mars

73 followed by its decomposition in acidic environments during Hesperian epoch (Fairén et al.,

74 2004; Bibring et al., 2006; Chevrier et al., 2007). Alternatively, abundant carbonates may never

75 have formed on early Mars because of short-term stability of liquid water and/or lack of dense

76 CO2 atmosphere during Noachian epoch (Bibring et al., 2006; Chevrier et al., 2007; Niles et al.,

77 2013; Edwards and Ehlmann, 2015; Zolotov and Mironenko, 2016). Acidic events could also

78 explain the apparent lack of abundant carbonate/phyllosilicate associations, because carbonate

79 do not form at acidic pH<6 conditions (Fairén et al., 2004; Fairén, 2013; Peretyazhko et

80 al., 2016; Zolotov and Mironenko, 2016). The carbonate/smectite mineralogical observation

81 might indicate that early Mars was not exclusively neutral-to-alkaline in pH but experienced

3

82 local and perhaps widespread acidic events. As a result, smectite formation on Mars could occur

83 not only under commonly expected neutral/alkaline conditions but also in acidic environments.

84 In terrestrial environments smectites formed under acidic conditions have been reported

85 in acidic saline lakes, seafloor hydrothermal vents and fumarolic areas (Haymon and Kastner,

86 1986; Story et al., 2010; Hynek et al., 2013). Limited laboratory observations have revealed

87 smectite formation through hydrothermal basalt alteration in acidic pH 3-6 environments (Berger

88 et al., 1987; Ghiara et al., 1993; Abdelouas et al., 1997; Dehouck et al., 2014). We have

89 demonstated formation of saponite through alteration of Mars-analogue glass-rich basalt

90 simulant in hydrothermal oxic and anoxic systems buffered by acetic acid (200 °C, pHRT ~4 pH

91 measured at room temperature, (Peretyazhko et al., 2016)).

92 The source of acidity on early Mars remains a subject for discussion. Two major acidity

93 sources have been proposed for early Mars: (1) Fe(II) oxidative hydrolysis (Tosca et al., 2008;

94 Hurowitz et al., 2010), and (2) volcanic release of sulfur dioxide, SO2 (Zolotov and Mironenko,

95 2007; Berger et al., 2009; Gaillard and Scaillet, 2009; Righter et al., 2009; Gaillard et al., 2013;

96 Zolotov and Mironenko, 2016). We have previously hypothesized that sulfuric acid (H2SO4)

97 produced by volcanic SO2 degassing on early Mars was the major source of acidity for the

98 alteration of basaltic materials and subsequent formation of smectite (Peretyazhko et al., 2016).

99 Our hypothesis was supported by modeling and mineralogical observations. For instance, recent

100 detection of smectite minerals co-existing with sulfates might indicate that basalt weathering and

101 smectite formation occurred under acidic conditions caused by sulfuric acid (Farrand et al., 2009;

102 Wray et al., 2011; Cavanagh et al., 2015; Flahaut et al., 2015; Rampe et al., 2017).

103 Thermodynamic modeling of basalt interaction with H2SO4-rich solutions revealed formation of

104 Al-rich phyllosilicates (kaolinite, montmorillonite) under mildly acidic conditions followed by

105 Mg/Fe smectites at higher pH (Zolotov and Mironenko, 2007; Zolotov and Mironenko, 2016)

4

106 and modelling of Mars-like aqueous systems predicted coexistence of smectite and sulfate

107 minerals under mildly acidic pH 4-6 conditions (Fairén, 2013).

108 Formation of smectite through basalt alteration under Mars-relevant acid sulfate

109 conditions has not been experimentally studied and the effect of pH and the nature of forming

110 phyllosilicate minerals remains unknown. The objective of this work was to investigate

111 formation of smectite through hydrothermal alteration of Mars-analogue basalt in the presence of

112 sulfuric acid of variable concentrations, to assess acidity sources and to determine the extent to

113 which H2SO4 is the source of acidity on early Mars.

114

115 2. MATERIALS AND METHODS

116

117 2.1. Synthesis and characterization of Adirondack basalt simulant

118 Synthetic basalt simulant of composition similar to that for Adirondack class rocks

119 analyzed by the at Crater (McSween et al., 2006) was used

120 in smectite formation studies (hereafter, denoted as Adirondack basalt simulant). The detailed

121 synthesis procedure of the Adirondack basalt simulant is reported by Peretyazhko et al., (2016).

122 Briefly, a powdered mixture of reagent-grade oxides and carbonates was melted at 1400 °C in an

123 Au-Pt alloy crucible for 3d under the oxygen fugacity IW+1 (IW = iron-wüstite buffer) and

124 quenched in water. The glassy product was then crushed, ground and sieved to <53 µm particle

125 diameter for characterization and smectite formation experiments. X-ray diffraction analysis

126 revealed that Adirondack simulant contained 66 wt% X-ray amorphous glass, 32 wt% olivine

127 and 2 wt% chromite quench crystals within the glass matrix (Peretyazhko et al., 2016).

128 Composition of the glass phase, olivine and chromite are summarized in Table EA-1.

129

5

130 2.2. Hydrothermal smectite formation experiments and characterization

131 Adirondack basalt simulant suspensions of 17 g/l (water to rock ratio = 60) were prepared

132 by mixing 250 mg simulant with 15 ml of solutions having variable H2SO4 concentrations (11.0

133 ± 0.1 mM, 13.6 ± 0.1 mM, 16.3 ± 0.1 mM, 21.9 ± 0.1 mM, 30.6 ± 0.3 mM and 42.5 ± 0.2 mM).

134 Initial sulfuric acid concentrations were measured as total sulfate by ion chromatography as

135 described below. Duplicate samples were placed in batch reactors (Teflon lined 23 ml Parr acid

136 digestion vessel) and incubated in an oven at 200 °C for 14d. The temperature of 200 °C was

137 chosen because hydrothermal conditions are favorable for smectite formation promoting

138 breaking of chemical bonds and rapid basalt alteration (Berger et al., 2014). However, smectite

139 formation through alteration of has been shown to occur at lower temperatures (Seyfried

140 et al., 1978; McMurtry and Yeh, 1981; Kloprogge et al., 1999; Hynek et al., 2013). Sample

141 preparation and hydrothermal incubation were performed under ambient atmosphere (oxidizing

142 conditions). Our previous study (Peretyazhko et al., 2016) demonstrated that smectite formed

143 through alteration of Adirondack basalt simulant under oxidizing and reducing acidic

144 hydrothermal conditions and that oxidizing conditions only affected Fe oxidation state in

145 smectite.

146 Suspension pH was measured immediately after mixing Adirondack basalt simulant with

147 H2SO4 (pH0d) and after opening the reactors at the end of 14d incubation (pH14d) with AccupHast

148 pH electrode using Orion Star A329 meter. Both measurements were performed at room

149 temperature (RT) of ~25 ºC. After pH measurement, solution was collected in a syringe and

150 passed through a 0.2 µm syringe filter (PVDF, Fisherbrand). An aliquot of each filtered solution

151 was acidified with ultra-pure HNO3 (Fisher) for dissolved Ca, Mg, Na, Si, Al and Fe analysis by

152 Inductively Coupled Plasma Mass Spectrometry (ICP-MS, Element XR Thermo Scientific) with

153 argon as the carrier gas. The remaining solution was analyzed for sulfate by ion chromatography

6

154 using Dionex ICS-2000 equipped with Dionex IonPac AS18 column (4 x 250 mm), Dionex EGC

155 III KOH eluent and a suppressed conductivity detector Dionex AERS 500 4 mm with a 20 µm

156 injection volume.

157 Separated solids were dried in an oven at 40 °C for ~1h prior to analysis. X-ray

158 diffraction (XRD) patterns were recorded using a Panalytical X’Pert Pro with Co Kα radiation.

159 Samples were analyzed at 45 kV/40 mA with a 0.02 2 step for 1 min. Programmable divergent

160 and antiscatter slits were set to 1/4 with manual 1/2 antiscatter slit. The instrument was

161 operated under ambient conditions and calibrated with novaculite (quartz) standard (Gemdugout,

162 State College, PA). Standard powder mounts and low background silicon holders were used for

163 analyses of bulk samples and clay size fractions, respectively. Clay size fractions were separated

164 by ultrasonic dispersion (Kachanoski et al., 1988). Smectite-containing suspensions were

165 ultrasonically treated with a probe-type sonicator (Soniprep 150) for 10s. The suspension was

166 then allowed to sediment by gravity for ~1hr and supernatant containing the clay size fraction

167 was collected. The procedure was repeated until supernatant solutions became clear and the

168 collected clay fraction was separated by centrifugation. The clay fraction slurry was spread over

169 the low background holder with a pipette and dried at RT prior to analysis. Clay fractions were

170 treated with glycerol and KCl in order to confirm formation of smectite (Whittig and Allardice,

171 1986).

172 Visible and near-infrared reflectance (VNIR) spectra (0.35-2.5 µm) were measured with

173 an Analytical Spectral Devices FieldSpec3 fiber-optic based spectrometer with a reflectance

174 probe attachment (internal light source). Spectra were collected in absolute reflectance mode

175 using a Spectralon standard. Continuum removed spectra were obtained using ENVI software

176 (ENvironment for Visualizing Images, Harris Geospatial Solutions).

7

177 Mössbauer analyses were performed at RT under ambient conditions using MIMOS-II

178 instruments that are the laboratory equivalent of the instruments onboard the Mars Exploration

179 Rovers [SPESI, Inc., (Klingelhoefer et al., 2003)]. Velocity calibration was carried out with the

180 program MERView (Agresti et al., 2007) using the MIMOS-II differential signal and the

181 spectrum for metallic Fe foil acquired at RT. Mössbauer parameters [center shift (CS) with

182 respect to metallic Fe foil at RT, quadrupole splitting (QS), magnetic hyperfine field strength

183 (Bhf), and subspectral areas (A)] were obtained with a least-squares fitting procedure using

184 MERFit (Agresti and Gerakines, 2009). Reported values for A were converted to % Fe using the

185 f-factor ratio, fFe(III)/f Fe(II) = 1.21 (Morris et al., 1995). Uncertainty for CS, QS, and full width at

186 half maximum (FWHM) was ± 0.02 mm/s and uncertainty for A was ± 2% absolute. The

187 following constraints were used in the least squares fitting procedure. (1) For each doublet, the

188 component peaks were constrained to have equal widths and areas. (2) Because of the low

189 observed intensity of hematite sextets, values for CS and QS were fixed to literature values

190 (e.g.,(Morris et al., 1985)), peak areas were constrained to 3 : x : 1 : 1 : x : 3, where x is variable

191 and FWHM decreased symmetrically toward zero velocity with skewed Lorentzian lineshapes.

192 Four representative samples were analyzed by Mössbauer spectroscopy covering the whole range

193 of the final pH14d values.

194 Electron microprobe analyses of polished sections of smectite-containing Adirondack

195 basalt simulant were analyzed for elemental composition using a CAMECA SX-100 electron

196 microprobe (Figure EA-1). Backscattered electron imaging and analyses were collected at an

197 accelerating voltage of 15 kV and probe current of 20 nA with a spot size of 1 µm. Element

198 concentrations were calibrated using University of Tokyo Oxides standards. All measurements

199 were made under 20s peak counting and 10s background counting. Because of high water

200 content in smectite, total elemental content was lower than 100% in all analyzed samples. Data

8

201 with totals ≥74 wt% were used for smectite formula calculation. Residual sulfur and

202 phosphorous were considered as impurities and subtracted from the totals in formula

203 calculations. All Si atoms were assigned to tetrahedral sites and the remaining tetrahedral sites

204 were filled with Al (Al = 8-Si). The remaining Al atoms, Mg, Fe and minor Ti, Ni, Cr and Mn

205 were assigned to the octahedral sites, while Na, Ca and K were assigned to interlayer sites (Bain

206 and , 1987).

207

208 2.3. Calculations of sulfur degassing on Mars

209 Crust production model and the amount of sulfur degassing from magma were used to

210 determine the release of SO2 during Noachian/Early Hesperian epoch on Mars and then to obtain

+ 211 the total amount of acidity (total amount of protons, H ) produced from the oxidation of SO2 to

212 H2SO4.

213 First, a crust production model described by Hirschmann and Withers (2008) was used to

214 calculate crust accumulation from early Noachian (~4.5 Ga) to late (0 Ga). The

215 model is based on assumptions that crust formation is proportional to the heat flow and that total

216 crust production during 4.5 Ga is 654·106 km3 as estimated by ( and Schneid, 1991). The

217 model is described by the equation:

218 (1)

219 where variables are V (crust production) and t (time) and constants are a = 9.277·105 km3, b =

220 2.385·105 km3, c = 6·10-3 Ma-1, d = 3.98·10-4 Ma-1 (Hirschmann and Withers, 2008). The

221 calculated total crust production is shown in Figure EA-2. The amount of sulfur released from

222 magma was then calculated as

223 Total S= Smagma ρ V (2)

9

3 224 where is an average basalt density of 2800 kg/m and Smagma is the amount of sulfur degassed

225 from magma of 2400 mg/kg (Righter et al., 2009).

226

227 3. RESULTS

228

229 3.1. pH before and after incubation

230 Different initial H2SO4 concentrations led to variable degrees of basalt neutralization

231 after 14d incubation. The average initial pH for all experiments was acidic (pH0d ≤ 2), and the

232 final pH14d varied over a wide range including acidic (pH14d 1.8 and 2.2), moderately acidic

233 (pH14d 3.1 and 3.7), neutral (pH14d 7.2) and slightly alkaline (pH14d 8.4) (Figure 1).

234 3.2. Mineralogical characterization

235 3.2.1. X-ray diffraction

236 Powder XRD analyses of reaction products showed increasing phyllosilicate formation at

237 pH14d ≥3 on the basis of increasing intensity of the 001 diffraction peak at ~15 Å while no

238 phyllosilicate formation occurred at pH14d 1.8 and 2.2 (Figure 2a, Table 1). After separation of

239 the clay size fraction, the phyllosilicate was identified as smectite based on expansion of the 001

240 peak to 17.4-18 Å after glycerol treatment, its collapse to 12.8 Å at room temperature upon KCl

241 addition, and its additional collapse to 10.1 Å after heating at 550 °C (Figure EA-3). The

242 positions of 02l and 060 d-spacings allowed distinguishing dioctahedral from trioctahedral

243 smectite (Moore and , 1997; Vaniman et al., 2014). The 02l and 060 d-spacings

244 revealed that trioctahedral smectite formed at pH14d 3.7, 7.2 and 8.4 and the peak positions

245 (Figure 2b, Table 2) were within the range reported for saponite (02l: 4.52-4.64 Å and 060: 1.53-

246 1.56 Å (Treiman et al., 2014; Chemtob et al., 2015; Peretyazhko et al., 2016)). The 02l and 060

247 d-spacings at 4.48 Å and 1.50 Å, respectively, indicated formation of dioctahedral smectite at

10

248 pH14d 3.1 (Figure 2b, Table 2). The peak positions were in agreement with d-spacing reported for

249 montmorillonite (02l: 4.47 Å and 060: 1.492-1.504 Å, (Moore and Reynolds, 1997; Vaniman et

250 al., 2014)). However, formation of montmorillonite could not be confirmed with VNIR and

251 electron microprobe analyses (sections 3.2.3 and 3.2.4) and therefore phyllosilicate formed in the

252 pH14d 3.1 sample is referred as dioctahedral smectite. Formation of smectite in all samples was

253 accompanied by decrease in intensity of the broad basaltic glass peak between 20° and 40° 2θ

254 with respect to the unaltered material (Figure 2a) suggesting glass as the smectite progenitor.

255 Anhydrite (CaSO4) was found to precipitate at all values of pH14d and natroalunite

256 (NaAl3(SO4)2(OH)6) was detected only at pH14d 1.8 (Figure 2a, Table 1). A weak diffraction

257 peak of hematite (α-Fe2O3) was observed in the pH14d 1.8 sample (Figure 2a, Table 1), and the

258 presence of hematite was confirmed by Mössbauer spectroscopy (Section 3.2.2). Hematite was

259 not detected in any other samples by XRD, but Mössbauer analysis showed that hematite was

260 present in the pH14d range from 1.8 to 8.4 (Section 3.2.2). Chromite, present in the unaltered

261 material, did not dissolve during incubation while olivine completely dissolved in the pH14d 1.8

262 and 2.2 samples as evident from the disappearance of its sharp diffraction peaks (Figure 2a,

263 Table 1).

264 3.2.2. Mössbauer spectroscopy

265 Mössbauer analysis of unaltered Adirondack basalt simulant showed that Fe was

266 associated with olivine and glass phases, 19 ± 2 and 81 ± 2% respectively and was mostly Fe(II)

267 (Table EA-2, (Peretyazhko et al., 2016)).

268 Analysis of altered Adirondack basalt simulant revealed formation of smectite through

269 alteration of the glass phase. The decrease in the concentration of total Fe associated with the

270 glass phase (39 ± 2% pH14d 3.1, 43 ± 2% pH14d 3.7 and 54 ± 2% pH14d 8.4, Figure 3a-c, Table

271 EA-2) compared to the unaltered simulant, was equal within error to amount of total Fe

11

272 associated with the smectite formed after 14d incubation (38 ± 2% pH14d 3.1, 41 ± 2% pH14d 3.7

273 and 56 ± 2% pH14d 8.4, Figure 3a-c, Table EA-2). Smectite spectra were fit with one Fe(II)

274 doublet (CS = 1.04 -1.09 mm/s and QS = 2.42 -2.66 mm/s) and one Fe(III) doublet (CS = 0.41 -

275 0.48 mm/s and QS = 0.54 - 0.70 mm/s QS, Table EA-2) and assigned to Fe(II) and Fe(III) in

276 octahedral coordination (Treiman et al., 2014). The smectite was oxidized with Fe(III)/Fe equal

277 to 0.61, 0.73 and 0.82 at pH14d 3.1, 3.7 and 8.4, respectively.

278 The most acidic pH14d 1.8 sample was also characterized by Fe(II) and Fe(III) doublets

279 with parameters similar to those assigned to smectite in the other samples (Fe(II) doublet: CS =

280 1.03 mm/s and QS = 2.60 mm/s and Fe(III) doublet: CS = 0.46 mm/s and QS = 0.62 mm/s,

281 Table EA-2). However, smectite did not form in this sample within XRD detection limits (Table

282 1). The Fe might be associated with an altered glass layer which forms during first stage of acidic

283 basalt alteration (Berger et al., 1987; Oelkers, 2001) and/or with an incipient phyllosilicate with

284 domains too small to coherently scatter X-rays.

285 The concentration of total Fe associated with olivine progressively decreased with

286 decreasing pH14d, and no Fe in olivine was detected in the pH14d 1.8 sample as shown in Table

287 EA-2. Olivine dissolution by acid-sulfate solutions could contribute Fe to smectite precipitation.

288 However, the Mössbauer evidence is that Fe(II) from olivine dissolution was oxidized and

289 precipitated as hematite, because the decrease in concentration of Fe(II) associated with olivine

290 was associated with a concomitant increase in the concentration of Fe(III) associated with

291 hematite (Figure 3, Table EA-2). Iron content in hematite (15 ± 2% pH14d 1.8, 12 ± 2% pH14d

292 3.1, 8 ± 2% pH14d 3.7 and 2 ± 2% pH14d 8.4, Figure 3, Table EA-2), was equal within an error of

293 fitting to decrease in olivine Fe content with respect to the unaltered Adirondack basalt simulant

294 (19 ± 2% pH14d 1.8, 11 ± 3% pH14d 3.1, 6 ± 3% pH14d 3.7 and 4 ± 3% pH14d 8.4, Figure 3, Table

12

295 EA-2). The hematite peaks were broad and skewed toward zero velocity, implying poor

296 crystallinity (small particle diameter) and potentially Al substitution for Fe(III).

297 3.2.3. Smectite composition and structural formula

298 The structural formula calculated from electron microprobe analysis of smectite-containing

299 particles (Table EA-3, Figure EA-1) showed that smectite contained Si and Al in tetrahedral

300 layers, Al, Mg, Fe with traces of Ti, Mn, Ni and Cr in octahedral layers, and Na, Ca and K

301 occupied interlayer sites. Given the high uncertainties in the measured data due to high water

302 content in smectite, the results of electron microprobe analysis did not allow to accurately

303 quantify smectite composition. Electron microprobe analysis indicated that smectite enriched in

304 Fe and Mg formed at pH14d 3.7, 7.2 and 8.4 while smectite equally enriched in Al, Mg and Fe

305 formed at pH14d 3.1 (Table EA-3). The average total number of octahedral cations in a unit cell

306 was close to 6 in the pH14d 3.7, 7.2 and 8.4 samples (Table EA-3) and was consistent with

307 trioctahedral saponite. The total number of octahedral cations was close to 5 in the pH14d 3.1

308 sample (Table EA-3) which might indicate formation of smectite containing dioctahedral and

309 trioctahedral domains.

310 3.2.4. Visible and near-infrared spectroscopy

311 VNIR spectral features between ~1.2 and 2.5 µm (Figure 4, Table 2) for saponite formed

312 at pH14d 3.7, 7.2, and 8.4 were assigned to OH in H2O and M-OH (~1.4 µm), to interlayer and

313 adsorbed H2O (~1.9 µm), and to trioctahedral M3-OH and dioctahedral M2-OH functional groups

314 (2.2-2.5 µm), where M is any combination of Al, Mg, Fe(II) and Fe(III) as octahedral cations.

315 The band centered between 2.30 and 2.32 μm was assigned to (Mg, Fe(II), Fe(III))3OH

316 combination stretching and bending vibration (Neumann et al., 2011; Treiman et al., 2014). (Mg,

317 Fe(II), Fe(III))3OH band center position was shifted to shorter wavelength in the pH14d 8.4

318 sample with respect to the pH14d 3.7 and 7.2 samples (Figure 4b, Table 2). The band position

13

319 shift could be due to increase in degree of Fe oxidation in saponite from Fe(III)/Fe = 0.73 at

320 pH14d 3.7 to Fe(III)/Fe = 0.82 at pH14d 8.4, (Table EA-2, (Peretyazhko et al., 2016)). The

321 position variability for the ~2.31 μm band might also result from heterogeneous distributions of

322 cations on octahedral sites [e.g., (Fe(III)2Mg)-OH versus (Fe(III)Mg2)-OH]. Saponite observed at

323 pH14d 7.2 and 8.4 had a weaker band near 2.40 μm band commonly observed in Fe/Mg smectites

324 (e.g., Carter et al., 2015; Chemtob et al., 2015). The exact band assignment is uncertain but

325 might be due to (Mg, Fe(II), Fe(III))3OH combination stretching and translation vibration (Frost

326 et al., 2002; Chemtob et al., 2015). The spectral feature near 2.24 μm was assigned to Al(Fe(III),

327 Fe(II), Mg)OH (Madejová et al., 2011; Chemtob et al., 2015) and indicated the presence of

328 octahedral Al, possibly present as dioctahedral domains as previously observed in synthetic

329 saponite (Chemtob et al., 2015; Peretyazhko et al., 2016). Spectral properties of saponite

330 synthesized at pH14d 3.7, 7.2 and 8.4 were similar to the properties reported for smectite on Mars

331 by VNIR spectrometers. The bands near 2.30 and 2.40 µm has been reported in Fe/Mg-smectite

332 identified on Mars while the band at 2.24 µm is present in about 10% of Fe/Mg-phyllosilicate

333 spectra (Carter et al., 2013), including nontronite (Bishop et al., 2013) but not saponite.

334 The spectrum of the pH14d 3.1 sample containing dioctahedral smectite was similar to the

335 spectra of the untreated Adirondack basalt simulant and pH14d 1.8 and 2.2 samples (Figure 4b).

336 All these samples had a broad band centered on 2.21-2.22 µm likely resulting from vibration of

337 Si-OH associated with the unaltered glass phase. The diagnostic band of montmorillonite at 2.20

338 µm was not distinguishable from the basalt band because of relatively low smectite abundance at

339 pH14d 3.1 (indicative by low intensity 001 diffraction peak in the sample, Figure 2a).

340 Hematite presence was evident from the ferric absorption edge between ~0.50 um and

341 ~0.75 um (Figure 4a) corresponding to the onset of strong absorption and a relative reflectivity

342 maximum for hematite (Morris et al., 1985). The VNIR manifestation of the hematite was most

14

343 noticeable for the pH14d 1.8 sample which had the highest hematite content (15% of total Fe,

344 Table EA-2)

345 3.3. Characterization of aqueous phase

346 Mineralogical transformations of Adirondack basalt simulant to smectite, anhydrite,

347 hematite and natroalunite (Table 1) resulted in release of variable proportions of Si, Mg, Na and

348 Ca. Complete dissolution of olivine in the pH14d 1.8 and 2.2 samples and subsequent release of

349 Mg and Si into solution (Figure 5a) were lower than the expected Si and Mg concentrations if all

350 olivine was dissolved (32 ± 3 mM Si and 42 ± 4 mM Mg based on the total SiO2 and MgO

351 contents in olivine, Table EA-1). The results suggest that Si and Mg dissolved from olivine were

352 partially incorporated into an altered glass phase under acidic conditions (Berger et al., 1987).

353 Smectite formation in the pH14d range from 3.1 to 8.4 (Figure 2a) led to decrease in Si

354 and Mg dissolution (Figure 5a) where the Mg dissolution likely controlled the nature of the

355 forming smectite (Meunier, 2005). Dioctahedral smectite was only detected in the pH14d

356 3.1 sample in which nearly 50% of the total Mg was leached into solution (calculations are based

357 on dissolved Mg concentrations in Figure 5a and total MgO in the glass phase, Table EA-1)

358 while formation of trioctahedral saponite (pH14d 3.7, 7.2 and 8.4) was accompanied by lower or

359 no Mg release into solution.

360 Dissolved Na was affected by precipitation of natroalunite and smectite formation.

361 Dissolved Na concentration was low and did not change in the pH14d range from 1.8 to 3.1 then

362 increased as pH shifted from acidic to slightly alkaline (Figure 5b). Low release of Na from

363 basaltic glass at pH14d 1.8-3.1 could result from precipitation of natroalunite, although the

364 presence of this phase was only confirmed at pH14d 1.8 (Table 1). Dissolved Na increased in the

365 pH14d range from 3.1 to 8.4 (Figure 5b) due to Na loss from glass phase during alteration of

366 Adirondack basalt simulant into smectite (Peretyazhko et al., 2016). Similar to Na, Ca was

15

367 expected to increase in solution with increasing basalt alteration as pH rose (Berger et al., 1987).

368 However, dissolved Ca concentration did not change (Figure 5b), and such non-pH-dependent

369 Ca release, together with the observed decrease in sulfate concentration (Figure 5c), resulted

370 from anhydrite precipitation (Figure 2a, Table 1) which is favorable at elevated temperatures

371 (Bishop et al., 2014).

372 Solution concentrations of Al and Fe were low for all samples resulting from their

373 incorporation into secondary phases (Figure EA-4, Tables 1 and 2). Dissolved Al did not exceed

374 0.04 mM (Figure EA-4a) as a result of precipitation of natroalunite at pH14d 1.8, low release from

375 glass alteration and incorporation into smectite at higher pH values (Tables 2 and EA-3) and

376 potentially in hematite. Dissolved Fe was <0.15 mM (Figure EA-4b) in all samples due to

377 precipitation of hematite accompanying olivine dissolution (Figure 3), low release from altering

378 glass phase and smectite formation (Tables 2 and EA-3).

379

380 4. DISCUSSION

381

382 The formation of smectite in the Noachian and early Hesperian has been attributed to

383 aqueous alteration of basaltic materials under neutral to alkaline pH aqueous conditions (Bibring

384 et al., 2006; Chevrier et al., 2007; Ehlmann et al., 2009; Bishop et al., 2013; Bristow et al., 2015;

385 Flahaut et al., 2015; Jain and Chauhan, 2015). However, the results of our experiments show that

386 formation of smectite from Mars-analogue basalt simulant is possible in much more acidic

387 environments. We found that gradual acid neutralization during acid-sulfate alteration of the

388 basalt simulant under hydrothermal conditions led to smectite formation at pH14d ≥ 3. The nature

389 of the forming smectite mineral varied with the solution final pH. Dioctahedral smectite equally

16

390 enriched in Al, Fe and Mg formed at pH14d ~3 followed by saponite enriched in Fe and Mg at

391 pH14d close to 4 or higher.

392 Our experimental results demonstrate that smectite formation through basalt alteration on

393 Mars was not limited to neutral/alkaline pH conditions but could partially occur under acidic

394 conditions once solution pH was close to 3 or higher. Sources of acidity on early Mars and the

395 persistence of acidic aqueous environments have been controversial. The basaltic crust should

396 provide enormous acid neutralizing potential (Niles and Michalski, 2009) suggesting that

397 aqueous systems on Mars should be largely alkaline. However, as noted above, absence of large

398 scale carbonate deposits shows an evidence for acidic conditions even in heavily altered regions

399 of the martian crust (Fairén et al., 2004). The principal acidity sources proposed for early Mars

400 are Fe(II) oxidative hydrolysis (Tosca et al., 2008; Hurowitz et al., 2010) and SO2 degassing

401 (Berger et al., 2009; Gaillard and Scaillet, 2009; Righter et al., 2009; Gaillard et al., 2013). We

402 argue here that SO2 degassing was the major source of acidity on early Mars generating H2SO4

403 sufficient for acidic phyllosilicate formation through basalt alteration and Fe(II) oxidative

404 hydrolysis (i.e., Fe(III) hydrolysis) was likely not the source of wide-spread acidity on Mars.

405 4.1. Acidity sources on early Mars: Fe(II) oxidative hydrolysis

406 Iron(III) hydrolysis was first suggested as the source of acidity at

407 (Hurowitz et al., 2010) and was further generalized as a possible widespread acidity source on

408 early Mars (e.g., (Ehlmann et al., 2011; Bowen et al., 2012; Horgan and Bell III, 2012;

409 McLennan, 2012; Marlow et al., 2014; Kaplan et al., 2016)). The Fe(II) oxidative hydrolysis

410 model proposed that upwelling of anoxic Fe(II)-containing groundwater generated from basalt

411 dissolution under anoxic circum-neutral pH conditions could lead to surface oxidation of Fe(II)

412 to Fe(III) followed by Fe(III) hydrolysis, release of H+ and precipitation of Fe(III) secondary

17

413 minerals (Hurowitz et al., 2010). The latter include nanophase iron(III) (hydr)oxides (interpreted

414 as schwertmannite (FeO(OH)0.75(SO4)0.125) (Hurowitz et al., 2010)), jarosite

415 ((K,Na)Fe3(SO4)2(OH)6) and hematite (α-Fe2O3) detected by Mössbauer spectroscopy at

416 Meridiani Planum (Morris et al., 2006). The reactions below summarize the mechanism

417 proposed for acidity generation following the sequence developed by Hurowitz et al., (2010)

418 with the exception that eq. 3 is included for basalt dissolution (eq. 3 written for olivine Fe(II)-end

419 member fayalite (Fe2SiO4)):

420 + 2+ 421 0.5Fe2SiO4 + 2H → Fe + 0.5SiO2 + H2O or (3a)

2+ - 422 0.5Fe2SiO4 + H2O → Fe + 0.5SiO2 + 2OH (3b)

2+ + 3+ 423 Fe + 0.25O2 + H → Fe +0.5H2O (4)

3+ 2- + 424 Fe + 0.125SO4 + 1.75H2O → FeO(OH)0.75(SO4)0.125 + 2.75H (5)

+ 2- + + 425 FeO(OH)0.75(SO4)0.125 + 0.25H2O + 0.75H + 0.5417SO4 + 0.33(K or Na ) →

426 0.33(K,Na)Fe3(SO4)2(OH)6 (6)

+ 2- 427 FeO(OH)0.75(SO4)0.125 + 0.25H2O → FeOOH + 0.25H + 0.125SO4 (7)

428 FeOOH → 0.5Fe2O3 + 0.5H2O (8)

429 While groundwater interaction with basalt was suggested to be a source of dissolved

430 Fe(II), basalt dissolution (eq. 3) was not included into Fe(II) oxidative hydrolysis model

431 (Hurowitz et al., 2010). As a result, the model predicted that sufficient amount of H+ was

432 produced through Fe(III) hydrolysis to develop acidic conditions (Hurowitz et al., 2010). If

433 basalt dissolution reactions are added to the model, then no acidity would be generated because

434 Fe(II) release is accompanied by OH- release (e.g., eq. 3b, Text EA-1). While our calculations

435 only account for Fe, the release of Mg, Ca, and Na from the basalts will also be accompanied by

436 release of OH- which will contribute to generation of additional alkalinity.

18

437 The OH- generated from dissolution of basalt (eq. 3) could be neutralized by H+ produced

438 through oxidative dissolution of Fe(II) sulfides present in basalt. However, the produced H+ will

439 not be sufficient to neutralize all released OH- on the basis of current estimates of low

440 abundances of Fe(II) sulfides in martian basalts (1600 mg/kg S in shergottite sulfides or 0.16

441 wt%, (Righter et al., 2009)). Alternatively, the OH- released from basalt together with Fe(II),

442 Mg, Ca and Na could be neutralized if the anoxic alkaline solutions encounter anoxic acidic

443 solutions. If all released OH- is neutralized, then Fe(II) oxidation followed by hydrolysis under

444 oxic conditions will lead to development of acidic conditions according to the Hurowitz et al.,

445 (2010) model (eqs. 4-8). However, such scenario will require presence of acidic conditions, for

446 instance H2SO4, before Fe(III) hydrolysis occurred.

447

448 4.2. Acidity sources on early Mars: SO2 degassing and H2SO4 formation

449 Magmatic sulfur degassing was the dominant source for sulfates deposited on the surface

450 of Mars (Righter et al., 2009), but was SO2 release on early Mars able to generate the amount of

451 H2SO4 sufficient for acidic phyllosilicate formation through alteration of basalt?

452 The calculated amount of sulfur release (eq. 2) during Noachian/early Hesperian epoch

453 was 2.5 x1021 g [~4.5 Ga - ~3.4 Ga; an intermediate value of 3.4 Ga was used for early

454 Hesperian/Late Hesperian boundary which is around 3.2 Ga - 3.6 Ga (Hartmann, 2005); Figure

20 + 455 6]. If all sulfur was released as SO2 and converted to H2SO4 then around 1.6x10 moles of H

456 was produced during dissociation of all acid in water. Formation of sulfuric acid might occur

457 through disproportionation of SO2 to H2SO4 and elemental S and S could be further transformed

458 into H2SO4 in the presence of oxidizing agent and water (Habashi and Bauer, 1966; Halevy et al.,

459 2007). It should be noted, however, an assumption that all degassed sulfur was transformed into

19

460 sulfuric acid likely overestimates the amount of produced acidity but our calculations allows to

461 potentially evaluate an upper limit of acidity that could have been generated on early Mars.

462 The protons generated from sulfuric acid are neutralized by Ca, Mg and Na in basalt via

463 proton exchange leading first to formation of an altered hydrated silica layer enriched in Fe and

464 Al (Berger et al., 1987; Peretyazhko et al., 2016) by schematic reactions:

465 Na-basalt + H+ = H-basalt + Na+ (9)

466 Ca-basalt + 2H+ = 2H-basalt + Ca2+ (10)

467 Mg-basalt + 2H+ = 2H-basalt + Mg2+ (11)

468 Crystallization of phyllosilicate from an altered layer does not substantially change pH as shown

469 in the previous studies of smectite formation through basalt alteration (Ghiara et al., 1993; De La

470 Fuente et al., 2002; Peretyazhko et al., 2016). Iron was not included into calculations because

471 overall process including dissolution of olivine followed by Fe(II) oxidation, Fe(III) hydrolysis

472 and hematite precipitation (Table 1) does not produce or consume H+ (Text EA-1).

+ 473 Approximately 9.4 molH+/kgbasalt H is required for complete neutralization of Na, Ca and Mg in

474 basalt of Adirondack composition (Text EA-2). The amount of H+ generated from volcanic

20 475 outgassing of SO2 (1.6x10 moles) during the Noachian and Hesperian would be sufficient for

476 neutralization of Ca, Mg and Na along with phyllosilicate formation in a 42 m thick basalt layer

477 over the entire (Text EA-3). Given that formation of smectite requires Mg to be present in

478 the altering basalt (Kloprogge et al., 1999), partial H+ neutralization by cations would be a more

479 realistic scenario on Mars. Noachian-aged terrains such as have been estimated to

480 contain 20-65% smectite (Poulet et al., 2008). Assuming that the smectite abundance in Mawrth

481 Vallis is representative of H+ neutralization and, consequently, degree of basalt alteration in

482 ancient martian terrains, the calculated average global depth of basaltic alteration would be

483 between 60 m (65% smectite content) to 200 m (20% smectite content) on the entire planet under

20

484 the constraint that alteration of 42 m thick basalt leads to 100% neutralization of Na, Ca and Mg.

485 The calculated depth range is consistent with phyllosilicate layer thickness of a few hundreds of

486 meters estimated in Noachian-aged terrains (Loizeau et al., 2012). Our acidity calculations and

487 experimental results, therefore, provide evidence that SO2 degassing would result in production

488 of large amounts of H+ and gradual neutralization of sulfuric acid during basalt alteration could

489 be an important mechanism of smectite formation on early Mars.

490

491 5. SIGNIFICANCE AND CONCLUSIONS

492

493 Our experiments demonstrated formation of smectite through hydrothermal acid-sulfate

494 alteration of Mars-analogue glass-rich basalt (66 wt% glass, 32 wt% olivine and 2 wt%

495 chromite). Basalt simulant suspensions were acidic with initial pH0d ≤ 2 while final pH14d varied

496 from acidic to mildly alkaline (pH14d 1.8-8.4) at the end of 14d incubation at 200 ºC. Alteration

497 of the glass phase of basalt simulant resulted in formation of dioctahedral smectite as pH14d

498 reached ~3 followed by trioctahedral saponite at pH14d close to 4 or higher. Reaction of

499 Adirondack basalt simulant with sulfuric acid also led to formation of anhydrite at all pH14d

500 values while natroalunite was detected at pH14d 1.8. Hematite precipitated as a result of partial or

501 complete oxidative dissolution of olivine in the pH14d range from 1.8 to 8.4.

502 Acidity on early Mars can be accounted for by volcanic outgassing of SO2 and formation

503 of sulfuric acid. Acid-sulfate conditions, suitable enough to form smectite and prevent carbonate

504 formation were likely sustained by low water/rock ratio and short duration of water-rock

505 interaction (Berger et al., 2009; Berger et al., 2014). Acidic smectite formation areas on early

506 Mars might be constrained to near-surface hydrothermal areas with magmatic outgassing

507 resulted from volcanic or impact activity (Solomon et al., 2005; Hynek et al., 2013). In these

21

508 systems only small portions of crust are exposed to acidity generated by H2SO4 formed from

509 degassed SO2. Short-term aquatic systems analogous to ephemeral acid saline lakes in Western

510 Australia might also allow development of acidic conditions (Story et al., 2010; Bowen et al.,

511 2012). Although there may also be acidity generated by the Fe(II) oxidative hydrolysis in local

512 environments, our calculations indicate that this mechanism is not a plausible large-scale acidity

513 source because of the alkalinity that is generated during basalt dissolution.

514 Formation of phyllosilicate by sulfuric acid alteration of basaltic materials on Mars would

515 likely be accompanied by precipitation of sulfate minerals including anhydrite. Although

516 anhydrite cannot be detected by the same remote sensing measurements used to detect

517 phyllosilicates because of the absence of diagnostic H2O spectral features in the VNIR, anhydrite

518 might coprecipitate with phyllosilicate in deposits formed under hydrothermal sulfuric acid

519 environments (>60 °C (Bishop et al., 2014)). Acid-sulfate smectite formation might also occur in

520 younger terrains. Phyllosilicates (e.g., smectite) co-existing with anhydrite in crater may

521 have formed along with the anhydrite under acid sulfate conditions (Rampe et al., 2017). The

522 hypothesis is supported by occurrence of jarosite in these mudstones in Gale crater indicative of

523 acidic environments (Rampe et al., 2017). Fe/Mg-rich phyllosilicates interbedded with sulfate

524 minerals in Meridiani Planum have been proposed to form prior to sulfates in moderate pH,

525 sulfate-free aqueous environments followed by sulfate precipitation from acidic surface and

526 groundwater (Flahaut et al., 2015). Our findings demonstrate that multiple events are not

527 required to form interbedded phyllosilicates and sulfates deposits. Acid-sulfate alteration can

528 explain martian mineralogical observations including mixed phyllosilicate/sulfate deposits and

529 lack of abundant carbonate deposits.

530

531

22

532 ACKNOWLEDGEMENTS

533

534 We thank Z. Peng for performing ICP-MS analysis. We thank Dr. Cuardos, Dr. Zolotov and two

535 anonymous reviewers for valuable suggestions and comments that help to improve the quality of

536 the manuscript. We thank the Associate Editor Dr. Catalano for handling the manuscript.This

537 work was supported by NASA’s Mars Fundamental Research Program grant # 11-MFRP11-

538 0090.

539

540

541

542

543

544

545

546

547

548

549

550

551

552

553

554

555

23

556 REFERENCES

557

558 Abdelouas A., Crovisier J.-L., Lutze W., Grambow B., Dran J.-C. and Müller R. (1997) Surface

559 layers on a borosilicate nuclear waste glass corroded in MgCl2 solution. J. Nucl. Mater. 240,

560 100-111.

561 Agresti D. G., Dyar M. D. and Schaefer M. W. (2007) Velocity scales for Mars Mössbauer data.

562 NASSAU 2006, 67-74.

563 Agresti D. G. and Gerakines P. A. (2009) Simultaneous fitting of Mars Mössbauer data. ICAME

564 2007, 1347-1354.

565 Bain D. C. and Smith B. F. L. (1987) Chemical analysis, In: A handbook of determinative

566 methods in clay mineralogy (ed. M. J. Wilson). Chapman and Hall, New York, pp. 248-275.

567 Berger G., Meunier A. and Beaufort D. (2014) Clay mineral formation on Mars: Chemical

568 constraints and possible contribution of basalt out-gassing. Planet. Space Sci. 95, 25-32.

569 Berger G., Schott J. and Loubet M. (1987) Fundamental processes controlling the first stage of

570 alteration of a basalt glass by seawater: an experimental study between 200 and 320 C. Earth.

571 Planet. Sci. Lett. 84, 431-445.

572 Berger G., Toplis M. J., Treguier E., d’Uston C. and Pinet P. (2009) Evidence in favor of small

573 amounts of ephemeral and transient water during alteration at Meridiani Planum, Mars. Am.

574 Mineral. 94, 1279-1282.

24

575 Bibring J.-P., Langevin Y., Mustard J. F., Poulet F., Arvidson R., Gendrin A., Gondet B.,

576 Mangold N., Pinet P. and Forget F. (2006) Global mineralogical and aqueous Mars history

577 derived from OMEGA/Mars Express data. Science 312, 400-404.

578 Bishop J. L., Lane M. D., Dyar M. D., King S. J., Brown A. J. and Swayze G. A. (2014) What

579 lurks in the martian rocks and soil? Investigations of sulfates, phosphates, and perchlorates.

580 Spectral properties of Ca-sulfates: gypsum, bassanite, and anhydrite. Am. Mineral. 99, 2105-

581 2115.

582 Bishop J. L., Loizeau D., McKeown N. K., Saper L., Dyar M. D., Des Marais D. J., Parente M.

583 and Murchie S. L. (2013) What the ancient phyllosilicates at Mawrth Vallis can tell us about

584 possible habitability on early Mars. Planet. Space Sci. 86, 130-149.

585 Bowen B. B., Benison K. C. and Story S. (2012) Early diagenesis by modern acid brines in

586 Western Australia and implications for the history of sedimentary modification on Mars. Mars

587 Sedimentology, SEPM Special Publication 102, 229-252.

588 Bristow T. F., Bish D. L., Vaniman D. T., Morris R. V., Blake D. F., Grotzinger J. P., Rampe E.

589 B., Crisp J. A., Achilles C. N. and Ming D. W. (2015) The origin and implications of clay

590 minerals from Yellowknife Bay, Gale crater, Mars. Am. Mineral. 100, 824-836.

591 Carter J., Poulet F., Bibring J. P., Mangold N. and Murchie S. (2013) Hydrous minerals on Mars

592 as seen by the CRISM and OMEGA imaging spectrometers: Updated global view. J. Geophys.

593 Res. 118, 831-858.

25

594 Cavanagh P. D., Bish D. L., Blake D. F., Vaniman D. T., Morris R. V., Ming D. W., Rampe E.

595 B., Achilles C. N., Chipera S. J., Treiman A. H., Downs R. T. and Morrison S. M. (2015)

596 Confidence Hills mineralogy and CheMin results from base of Mt. Sharp, Pahrump Hills, Gale

597 Crater, Mars LPSC.

598 Chemtob S. M., Nickerson R. D., Morris R. V., Agresti D. G. and Catalano J. G. (2015)

599 Synthesis and structural characterization of ferrous trioctahedral smectites: Implications for clay

600 mineral genesis and detectability on Mars. J.Geophys. Res. Planets. 120, 1119 -1140

601 Chevrier V., Poulet F. and Bibring J.-P. (2007) Early geochemical environment of Mars as

602 determined from thermodynamics of phyllosilicates. Nature 448, 60-63.

603 Craddock R. A. and Greeley R. (2009) Minimum estimates of the amount and timing of gases

604 released into the martian atmosphere from volcanic eruptions. Icarus 204, 512-526.

605 De La Fuente S., Cuadros J. and Linares J. (2002) Early stages of volcanic tuff alteration in

606 hydrothermal experiments: formation of mixed-layer illite-smectite. Clays Clay Miner. 50, 578-

607 590.

608 Dehouck E., Gaudin A., Mangold N., Lajaunie L., Dauzères A., Grauby O. and Le Menn E.

609 (2014) Weathering of olivine under CO2 atmosphere: a martian perspective. Geochim.

610 Cosmochim. Acta 135, 170-189.

611 Edwards C. S. and Ehlmann B. L. (2015) Carbon sequestration on Mars. Geo 43, 863-866.

26

612 Ehlmann B. L., Mustard J. F., Murchie S. L., Bibring J.-P., Meunier A., Fraeman A. A. and

613 Langevin Y. (2011) Subsurface water and clay mineral formation during the early history of

614 Mars. Nature 479, 53-60.

615 Ehlmann B. L., Mustard J. F., Murchie S. L., Poulet F., Bishop J. L., Brown A. J., Calvin W. M.,

616 R. N., Des Marais D. J. and Milliken R. E. (2008) Orbital identification of carbonate-

617 bearing rocks on Mars. Science 322, 1828-1832.

618 Ehlmann B. L., Mustard J. F., Swayze G. A., Clark R. N., Bishop J. L., Poulet F., Des Marais D.

619 J., Roach L. H., Milliken R. E. and Wray J. J. (2009) Identification of hydrated silicate minerals

620 on Mars using MRO‐CRISM: Geologic context near and implications for aqueous

621 alteration. J. Geophys. Res. Planets 114.

622 Fairén A. G. (2013) Coeval synthesis of cold aqueous mineralogies on Mars, In: Mars: evolution,

623 geology and exploration (ed. A. G. Fairén). Nova Science Publisher.

624 Fairén A. G., Fernández-Remolar D., Dohm J. M., Baker V. R. and Amils R. (2004) Inhibition

625 of carbonate synthesis in acidic oceans on early Mars. Nature 431, 423-426.

626 Farrand W. H., Glotch T. D., Rice J. W., Hurowitz J. A. and Swayze G. A. (2009) Discovery of

627 jarosite within the Mawrth Vallis region of Mars: Implications for the geologic history of the

628 region. Icarus 204, 478-488.

629 Flahaut J., Carter J., Poulet F., Bibring J.-P., van Westrenen W., G. and Murchie S.

630 (2015) Embedded clays and sulfates in Meridiani Planum, Mars. Icarus 248, 269-288.

27

631 Frost R. L., Kloprogge J. T. and Ding Z. (2002) Near-infrared spectroscopic study of nontronites

632 and ferruginous smectite. Spectrochim. Acta 58, 1657-1668.

633 Gaillard F., Michalski J., Berger G., McLennan S. M. and Scaillet B. (2013) Geochemical

634 reservoirs and timing of sulfur cycling on Mars. Space Sci. Rev. 174, 251-300.

635 Gaillard F. and Scaillet B. (2009) The sulfur content of volcanic gases on Mars. Earth. Planet.

636 Sci. Lett. 279, 34-43.

637 Ghiara M., Franco E., Petti C., Stanzione D. and Valentino G. (1993) Hydrothermal interaction

638 between basaltic glass, deionized water and seawater. Chem. Geol. 104, 125-138.

639 Greeley R. and Schneid B. (1991) Magma generation on Mars- Amounts, rates, and

640 comparisons with earth, moon, and Venus. Science 254, 996-998.

641 Habashi F. and Bauer E. (1966) Aqueous oxidation of elemental sulfur. Ind. Eng. Chem.

642 Fundam. 5, 469-471.

643 Halevy I., Zuber M. T. and Schrag D. P. (2007) A sulfur dioxide climate feedback on early

644 Mars. Science 318, 1903-1907.

645 Hartmann W. K. (2005) Martian cratering 8: Isochron refinement and the chronology of Mars.

646 Icarus 174, 294-320.

647 Haymon R. M. and Kastner M. (1986) The formation of high temperature clay minerals from

648 basalt alteration during hydrothermal discharge on the East Pacific Rise axis at 21 N. Geochim.

649 Cosmochim. Acta 50, 1933-1939.

28

650 Hirschmann M. M. and Withers A. C. (2008) Ventilation of CO2 from a reduced mantle and

651 consequences for the early Martian greenhouse. Earth. Planet. Sci. Lett. 270, 147-155.

652 Horgan B. and Bell III J. F. (2012) Widespread weathered glass on the surface of Mars. Geology

653 40, 391-394.

654 Hurowitz J. A., Fischer W. W., Tosca N. J. and Milliken R. E. (2010) Origin of acidic surface

655 waters and the evolution of atmospheric chemistry on early Mars. Nat. Geosci.3, 323-326.

656 Hynek B. M., McCollom T. M., Marcucci E. C., Brugman K. and Rogers K. L. (2013)

657 Assessment of environmental controls on acid‐sulfate alteration at active volcanoes in

658 Nicaragua: Applications to relic hydrothermal systems on Mars. J. Geophys. Res. Planets 118,

659 2083-2104.

660 Jain N. and Chauhan P. (2015) Study of phyllosilicates and carbonates from the Capri

661 region of on Mars based on Mars Reconnaissance Orbiter-Compact

662 Reconnaissance Imaging Spectrometer for Mars (MRO-CRISM) observations. Icarus 250, 7-17.

663 Kachanoski R., Voroney R. and Gregorich E. (1988) Ultrasonic dispersion of aggregates:

664 distribution of organic matter in size fractions. Can. J. Soil. Sci. 68, 395-403.

665 Kaplan H. H., Milliken R. E., Fernández-Remolar D., Amils R., Robertson K. and Knoll A. H.

666 (2016) Orbital evidence for clay and acidic sulfate assemblages on Mars based on mineralogical

667 analogs from Rio Tinto, Spain. Icarus 275, 45-64.

29

668 Klingelhoefer G., Morris R. V., Bernhardt B., Rodionov D., De Souza P., Squyres S., Foh J.,

669 Kankeleit E., Bonnes U. and Gellert R. (2003) Athena MIMOS II Mössbauer spectrometer

670 investigation. J. Geophys. Res. Planets 108.

671 Kloprogge J. T., Komarneni S. and Amonette J. E. (1999) Synthesis of smectite clay minerals: a

672 critical review. Clays Clay Miner. 47, 529-554.

673 Loizeau D., Werner S., Mangold N., Bibring J.-P. and Vago J. (2012) Chronology of deposition

674 and alteration in the Mawrth Vallis region, Mars. Planet. Space Sci. 72, 31-43.

675 Madejová J., Balan E. and Petit S. (2011) Application of vibrational spectroscopy to the

676 characterization of phyllosilicates and other industrial minerals. Advances in the

677 Characterization of Industrial Minerals: Notes in Mineralogy 9, 171-226.

678 Marlow J. J., LaRowe D. E., Ehlmann B. L., Amend J. P. and Orphan V. J. (2014) The potential

679 for biologically catalyzed anaerobic methane oxidation on ancient Mars. Astrobiology 14, 292-

680 307.

681 McLennan S. M. (2012) Geochemistry of sedimentary processes on Mars. Sediment. Geol.Mars

682 102, 119-138.

683 McMurtry G. M. and Yeh H.-W. (1981) Hydrothermal clay mineral formation of East Pacific

684 Rise and Bauer Basin sediments. Chem. Geol. 32, 189-205.

30

685 McSween H. Y., Wyatt M. B., Gellert R., Bell J., Morris R. V., Herkenhoff K. E., Crumpler L.

686 S., Milam K. A., Stockstill K. R. and Tornabene L. L. (2006) Characterization and petrologic

687 interpretation of olivine‐rich basalts at Gusev Crater, Mars. J. Geophys. Res. Planets 111.

688 Meunier A. (2005) Clays. Springer Science & Business Media.

689 Milliken R., Fischer W. and Hurowitz J. (2009) Missing salts on early Mars. Geophys. Res. Lett.

690 36.

691 Moore D. M. and Reynolds R. C. J. (1997) X-ray diffraction and the identification and analysis

692 of clay minerals. Oxford University Press, New York.

693 Morris R. V., Golden D., Bell J. F. and Lauer H. (1995) Hematite, , and phyllosilicates

694 on Mars: Implications from oxidized impact melt rocks from Manicouagan Crater, Quebec,

695 Canada. J. Geophys. Res. Planets 100, 5319-5328.

696 Morris R. V., Klingelhoefer G., Schröder C., Rodionov D. S., Yen A., Ming D. W., De Souza P.,

697 Wdowiak T., Fleischer I. and Gellert R. (2006) Mössbauer mineralogy of rock, soil, and dust at

698 Meridiani Planum, Mars: 's journey across sulfate‐rich outcrop, basaltic sand and

699 dust, and hematite lag deposits. J. Geophys. Res. Planets 111.

700 Morris R. V., Lauer H. V., Lawson C. A., Gibson E. K., Nace G. A. and Stewart C. (1985)

701 Spectral and other physicochemical properties of submicron powders of hematite (α‐Fe2O3),

702 maghemite (γ‐Fe2O3), magnetite (Fe3O4), goethite (α‐FeOOH), and lepidocrocite (γ‐FeOOH). J.

703 Geophys. Res. Solid Earth 90, 3126-3144.

31

704 Morris R. V., Ruff S. W., Gellert R., Ming D. W., Arvidson R. E., Clark B. C., Golden D.,

705 Siebach K., Klingelhöfer G. and Schröder C. (2010) Identification of carbonate-rich outcrops on

706 Mars by the Spirit rover. Science 329, 421-424.

707 Murchie S. L., Mustard J. F., Ehlmann B. L., Milliken R. E., Bishop J. L., McKeown N. K., Noe

708 Dobrea E. Z., Seelos F. P., Buczkowski D. L. and Wiseman S. M. (2009) A synthesis of Martian

709 aqueous mineralogy after 1 Mars year of observations from the Mars Reconnaissance Orbiter. J.

710 Geophys. Res. Planets 114.

711 Neumann A., Petit S. and Hofstetter T. B. (2011) Evaluation of redox-active iron sites in

712 smectites using middle and near infrared spectroscopy. Geochim. Cosmochim. Acta 75, 2336-

713 2355.

714 Niles P. B., Catling D. C., Berger G., Chassefière E., Ehlmann B. L., Michalski J. R., Morris R.,

715 Ruff S. W. and Sutter B. (2013) Geochemistry of : implications for climate

716 history and nature of aqueous environments. Space Scie. Rev. 174, 301-328.

717 Niles P. B. and Michalski J. (2009) Meridiani Planum sediments on Mars formed through

718 weathering in massive ice deposits. Nat. Geosci. 2, 215-220.

719 Oelkers E. H. (2001) General kinetic description of multioxide silicate mineral and glass

720 dissolution. Geochim. Cosmochim. Acta 65, 3703-3719.

721 Peretyazhko T., Sutter B., Morris R., Agresti D., Le L. and Ming D. (2016) Fe/Mg smectite

722 formation under acidic conditions on early Mars. Geochim. Cosmochim. Acta 173, 37-49.

32

723 Poulet F., Bibring J.-P., Mustard J., Gendrin A., Mangold N., Langevin Y., Arvidson R., Gondet

724 B., Gomez C. and Berthé M. (2005) Phyllosilicates on Mars and implications for early Martian

725 climate. Nature 438, 623-627.

726 Poulet F., Mangold N., Loizeau D., Bibring J.-P., Langevin Y., Michalski J. and Gondet B.

727 (2008) Abundance of minerals in the phyllosilicate-rich units on Mars. Astron.Astrophys. 487,

728 L41-L44.

729 Rampe E. B., Ming D. W., Blake D. F., Bristow T. F., Chipera S. J., Grotzinger J. P., Morris R.

730 V., Morrison S. M., Vaniman D. T., Yen A. S. and et al (2017) Mineralogy of an ancient

731 lacustrine mudstone succession from the Murray formation, Gale

732 crater, Mars. Earth. Planet. Sci. Lett. 471, 172-185.

733 Righter K., Pando K. and L. (2009) Experimental evidence for sulfur-rich martian

734 magmas: Implications for volcanism and surficial sulfur sources. Earth. Planet. Sci. Lett. 288,

735 235-243.

736 Seyfried W., Shanks W. and Dibble W. (1978) Clay mineral formation in DSDP Leg 34 basalt.

737 Earth. Planet. Sci. Lett. 41, 265-276.

738 Solomon S. C., Aharonson O., Aurnou J. M., Banerdt W. B., Carr M. H., Dombard A. J., Frey H.

739 V., Golombek M. P., Hauck S. A. and Head J. W. (2005) New perspectives on ancient Mars.

740 Science 307, 1214-1220.

741 Story S., Bowen B. B., Benison K. C. and Schulze D. G. (2010) Authigenic phyllosilicates in

742 modern acid saline lake sediments and implications for Mars. J. Geophys. Res. Planets 115.

33

743 Tosca N. J., McLennan S. M., Dyar M. D., Sklute E. C. and Michel F. M. (2008) Fe oxidation

744 processes at Meridiani Planum and implications for secondary Fe mineralogy on Mars. J.

745 Geophys. Res. Planets 113.

746 Treiman A. H., Morris R. V., Agresti D. G., T. G., Achilles C. N., Rampe E. B., Bristow T.

747 F., Blake D. F., Vaniman D. T. and Bish D. L. (2014) Ferrian saponite from the Santa Monica

748 Mountains (California, USA, Earth): Characterization as an analog for clay minerals on Mars

749 with application to Yellowknife Bay in Gale Crater. Am. Mineral. 99, 2234-2250.

750 Vaniman D., Bish D., Ming D., Bristow T., Morris R., Blake D., Chipera S., Morrison S.,

751 Treiman A. and Rampe E. (2014) Mineralogy of a mudstone at Yellowknife Bay, Gale crater,

752 Mars. Science 343, 1243480.

753 Whittig L. D. and Allardice W. R. (1986) X-ray diffraction techniques, In: Methods of soil

754 analysis. Part 1 Physical and mineralogical methods (ed. A. Klute). American Society of

755 Agronomy, Madison, pp. 331-363.

756 Wray J., Milliken R., Dundas C., Swayze G., Andrews‐Hanna J., Baldridge A., Chojnacki M.,

757 Bishop J., Ehlmann B. and Murchie S. L. (2011) crater and other possible

758 groundwater‐fed paleolakes of Terra Sirenum, Mars. J. Geophys. Res. Planets 116.

759 Wray J. J., Murchie S. L., Bishop J. L., Ehlmann B. L., Milliken R. E., Wilhelm M. B., Seelos K.

760 D. and Chojnacki M. (2016) Orbital evidence for more widespread carbonate‐bearing rocks on

761 Mars. J. Geophys. Res. Planets 121, 652-677.

34

762 Zolotov M. Y. and Mironenko M. V. (2007) Timing of acid weathering on Mars: A

763 kinetic‐thermodynamic assessment. J. Geophys. Res. Planets 112.

764 Zolotov M. Y. and Mironenko M. V. (2016) Chemical models for martian weathering profiles:

765 Insights into formation of layered phyllosilicate and sulfate deposits. Icarus 275, 203-220.

766

767

768

35

769 Table 1. Phases determined in the 14d-incubated Adirondack basalt simulants by XRD and

770 Mössbauer spectroscopy. Olivine and chromite are unaltered phases from the original basalt

771 simulant.

pH14d Crystalline phases 1.81 ± 0.01 anhydrite, natroalunite, hematitea, chromite 2.19 ± 0.01 anhydrite, chromite 3.12 ± 0.03 dioctahedral smectite, anhydrite, hematite, olivine, chromite 3.72 ± 0.34 saponite, anhydrite, hematite, olivine, chromite 7.17 ± 0.38 saponite, anhydrite, olivine, chromite 8.38 ± 0.20 saponite, anhydrite, hematite, olivine, chromite a 772 Hematite was detected by both XRD and Mössbauer in the pH14d 1.8 sample and by Mössbauer 773 spectroscopy alone in the other samples. Hematite was present within the pH14d range from 1.8 to 774 8.4 and likely precipitated in the pH14d 2.2 and 7.2 samples not analyzed by Mössbauer 775 spectroscopy. 776

777

778

779

780

781

782

783

784

785

786

787

788

36

789

790 Table 2. Smectite characterization data obtained by XRD and VNIR. pH14d Smectite XRD VNIR d-spacing, Å combination band position, µm 02l 060 Al(Fe(III), (Mg, Fe(II), (Mg, Fe(II), b b c Fe(II), Mg)OH Fe(III))3OH Fe(III))3OH dioctahedral 3.12 ± 0.03 4.48 1.50 nda nd nd smectite 3.72 ± 0.34 saponite 4.55 1.53 2.239 2.319 nd 7.17 ± 0.38 saponite 4.58 1.53 2.246 2.319 2.403 8.38 ± 0.20 saponite 4.57 1.53 2.243 2.300 2.408 791 a nd-not detected; b Combination stretching and bending vibration (Neumann et al., 2011; 792 Treiman et al., 2014); c Combination stretching and translation vibration (Frost et al., 2002; 793 Chemtob et al., 2015). 794

795

796

797

798

37

pH0d 8 pH14d

6

pH

4

2

10 20 30 40 Initial sulfate concentration, mM 799 800 801 Figure 1. Initial (pH0d) and final (pH14d) pH in Adirondack basalt simulant suspensions as a 802 function of initial sulfate concentrations. 803

38

804 805 Figure 2. (a) XRD for unaltered and 14d-incubated Adirondack basalt simulant samples with 806 different pH14d. S = smectite (001, 02l and 060 diffraction peaks are marked next to or above the 807 peaks), A = anhydrite, Ol = olivine, C = chromite, NA = natroalunite, H = hematite; (b) 02l and 808 060 diffraction bands for clay fractions separated from the pH14d 8.4 and 3.1 samples. The bands 809 were analyzed with a 0.02 2 step for 1 min and 0.02 2 step for 17 min in the pH14d 8.4 and 810 3.1 clay fractions, respectively.

39

811

812 Figure 3. Mössbauer spectra and fit subspectra for 14d-incubated (a) pH14d 8.4, (b) pH14d 3.7, (c) pH14d 3.1 and (d) pH14d 1.8 Adirondack 813 basalt simulant collected at RT [subspectral areas in % are for Ol = olivine; Total Gl = total glass, a sum of 2D1 and 2D2 doublets 814 (Table EA-2), Total Sme = total smectite, and Total Alt Gl = total altered glass, a sum of 3D2 and 2D4 doublets (Table EA-2)]. TC = 815 total counts, BC = baseline counts.

816 817 Figure 4. (a) VNIR reflectance spectra of unaltered and 14d-incubated Adirondack basalt 818 simulant samples with different pH14d. Bands at 1.4 and 1.9 µm in the saponite-containing pH14d 819 8.4, 7.2 and 3.7 samples are marked with dotted lines. (b) 2-2.5 µm range of continuum removed 820 reflectance spectra. Positions of M3-OH and M2-OH combination bands in the saponite- 821 containing samples are marked with dotted lines (M is any combination of Al, Mg, Fe(II) and 822 Fe(III) as octahedral cations). Characteristic Al2OH band of montmorillonite at 2.21 µm in the 823 pH14d 3.1 sample was not detected. 824 825

826 Figure 5. Concentrations of dissolved (a) Mg and Si, (b) Na and Ca and (c) sulfate in Adirondack 827 basalt simulant suspensions as a function of pH14d. 828

42

829 830 Figure 6. Sulfur degassed on Mars from 4.5 Ga to present calculated using the crust production 831 model (eq. 2, Hirschmann and Withers 2008) and amount of sulfur degassed from 1 kg magma 832 (2400 mg/kg, Righter et al., 2009). The calculated total degassed sulfur from 4.5 Ga to present 833 was 4.4·1021 g and within the range of estimates of total sulfur release on Mars: 2.2·1022 g 834 (McLennan, 2012), 5.4·1021 g (Gaillard and Scaillet, 2009), 1.7·1020 g (Craddock and Greeley, 835 2009) and 4.5·1019 g (Righter et al., 2009). Dashed line shows a boundary between early and late 836 Hesperian. 837

43